A Carbon-13 Nuclear Magnetic Resonance Study ... - ACS Publications

2118 J. Org. Chem.. G'ol. 42, No. 12, 1977. Hall and Wemple ... J ,6.6 Hz, 3 H, C-2 CHs); l,'iC NMR, see Table I; high-resolution MS, obsdhalcd mass (...
0 downloads 0 Views 743KB Size
2118

J . Org. Chem.. G'ol. 42, No. 12, 1977

Hall and Wemple

CHCI3)Hf0.38; IR (CC14)2778 (Bohlmann bands), 873 cm-l (3-furyl); 1H NMR 6 7.31 (m, 2 H, 3-fury1 n-H), 6.36 (m, 1 ,3-furyl P-H), 3.60 (dd, J = 10.4 and 3.0 Hz, 1 H, C-2 H), 2.78 (q'd, 1 H, C-6 H),1.05 (d, J ,6.6 Hz, 3 H, C-2 CHs); l,'iCNMR, see Table I; high-resolution MS, obsdhalcd mass (formula I 165.1151/165.1152 (CIOH~RNO). Fraction E4 was chromai;ographed on a 1 X 11cm column of neutral fractions and a 30-ml CHC13fraction (F4).Fraction F4 was chromatographed on a 0.7 X 2-cm column of neutral Alumina (activity 2) with 150 mL of CHzC12 (Gl) and two 75-mL portions of CHC13-CH2C12 (1:9) (G2 and G3). Fractions F3 (3 mg) and G3 (7 mg) were combined and consisted of pure 22: GLC (column E, 150 "C) Rt 6.06; TLC (Alumina, twice developetd, once with CHzCl2 and then CHC13) R f 0.31; IR (CCld) 2718 (very weak Bohlmann bands), 873 cm-l (3-furyl); 1H NMR 6 7.24-7.44 (m, 2 H, 3-fury1n-H), 6.36 (m, 1 H, 3-fury1O-H), 4.10 (t, J = 4 Hz, 1 H, C-2 H), 3.04 (q'd, J = 7.0 and 2.2 Hz, 1H, C-6 H), 1.11 (d, J = 6.0 Hz, 3 14, C-6 CH3); 13C NMR, see Table I; highresolution MS, obsd/calcd mass (formula) 165.1173/165.1152 (CioHdO). Registry No.-1, 61949-86-8; 2, 61949-87-9; 3, 110-93-0; 4, 53067-23-5;5,61900-43-4; 6,61900-44-5; 7,61900-45-6; 8,61900-46-7; 9, 61900-47-8; 13, 61900-48-9; 15, 61586-90-1; 17, 61900-30-9; 18, 61900-31-0; 19, 61900-32-1; 20, 61900-33-2; 21, 61900-34-3; 22, 61900-35-4; 23, 61900-36-5; 24, 61900-37-6; 25, 61900-38-7; 27, 61949-85-7; dimethyl carbonate, 616-38-6; hydroxylamine HCl, 5470-11-1; dimethyl sulfate, 77-78-1; 3-furoyl chloride, 26214-65-3; 5-aminopentanoic acid, 660-88-8; 5-(3-furamido)hexanoate, 61900-39-8; (-) -nuphenin e, 4850-01-5; (-)-anhydronupharamine, 4849-88-1.

References a n d Notes Support of this work by the National Institute of Health, U S . Public Health Service (Grant AI 10166) is gratefully acknowledged, (2) E. Maurer and G. Ohloff, Helv. Chim.Acta, 59, 1169 (1976). (1)

(3) R. T. LaLonde and C. F. Wong. Can. J. Chem., 53, 1714 (1976). (4) Previous syntheses of piperidine Nuphar alkaloids are: (*)-nupharamlne and (iF3-epinupharamine,J. Szychowski,J. T. Wrbbel. and A. Lenlewski, Bull. Acad. Pol. Sci. Ser. Sci. Chim., 5,385 (1974); Kmethylnupharamine, S. Matsutani. I. Kawasaki,and T. Kaneko, Bull. Chem. Soc.Jpn., 40,2629 (1967). (5) D. Varech, C. Ouannes, and J. Jacques, Bull. Soc. Chim. Fr., 1662 (1965). (6) F. E. LaForge. N. Green, and W. A. Gersdorff, J. Am. Chem. Soc., 70,3707 (1946). (7) T. Fukuyama, L. V. Dunkerton,M. Aratani, and Y. Kishi, J. Org. Chem., 40, 2011 (1975). (8) J. E. Oatis, Jr., unpublished observations. (9) R. T. LaLonde, A. El-Kafrawy, N. Muhammad, and J. E. Oatls, Jr., J. Org. Chem., in press. (10) Y. Arata, T. Ohashi, M. Yonemitsu, and S. Yasuda. Yakugaku Zasshi, 87, 1094 (1967). (11) (a)I. Kawasaki, S. Matsutani, and T. Kaneko, Bull. Chem. SOC. Jpn., 36, 1474 (1936); (b) R. T. Lalonde, E. Auer, C. F. Wong, and V. P. Muralidharan, J. Am. Chem. Soc., 93, 2501 (1971). (12) N. Oda and H. Koyama, J. Chem. Soc.8, 1450 (1970). (13) R. Barchet and T. P. Forrest, Tetrahedron Lett., 4229 (1965). (14) T. P. Forrest and S. Rav. Can. J. Chem.. 49. 1774 11971). (15) T. M. Moynehan, K. Schbfleld,R. A. Y. J&;and A.'R. K a k k y , J. Chem. Soc.. 2637 (1962). (16) J. B..Stothe;s, "Carbon-13NMR Spectroscopy". Academic Press, New York, N.Y., 1972, p 272. (17) All 3-fury1 carbon resonance values, except that for the nonprotonated carbon in 27, fall within the range of values for each of the four types of carbon atoms as found in: R. T. LaLonde. T. N. Donvito, and A. I-M. Tsai, Can. J. Chern., 53, 1714 (1975). The nonprotonated 3-fwyl carbon In 27 falls at 6 126.6 ppm, which is 0.5 ppm below the lower limit of the range for this type of carbon.The resonance values for the vinylmethyland oleflnlc carbons correspond to those given in: J. E. Stothers, ref 16, pp 72 and 82. (18) H. Booth and D. V. Griffiths, J. Chem. Soc.,Perkin Trans. 2, 843 (1973). (19) Our 13C NMR resonance assignmentsfor anhydronupharamineagree with those of another study S. Yasuda, M. Hanaoka,and Y. Arata, Chem. pherm. Bull., 24, 2521 (1976h which we became aware of after this paper had

been submitted for pubilcation.

A Carbon-13 Nuclear Magnetic Resonance Study of Thiol Esters Charles M. Hall* Research Laboratories, T h e Upjohn Company, Kalamazoo, Michigan 49001

James Wemple* Department of Chemistry and Chemical Engineering, University of Detroit, Detroit, Michigan 48221 Received December 14,1976 The 13C NMR chemical shifts for each of the carbons of a number of simple thiol esters have been measured in MezSO-d6 and CDC13 and the results related to known chemical properties of the thiol ester function. In general the chemical shift of the thiol ester carbonyl carbon occurs some 15-20 ppm further downfield than that found for all other carboxylic acid derivatives reported to date. The carbon a to the carbonyl function in thiol esters is also shifted downfield by about 10 ppm relative to the a carbon in acids, oxygen esters, or amides. The effect of carbon group and halogen substituents on thiol ester chemical shifts has been analyzed. A solvent study on &hydroxy thiol esters shows that the carbonyl carbon is shielded in MezSO-ds relative to CDC13, which may be attributed to intramolecular hydrogen bonding in the latter solvent. Carbon-13 chemical shift changes resulting from conversion of mercaptans to thiol ester derivatives indicate relatively little difference between S-tert -butyl and other types of S-alkyl thiol esters in contrast to results obtained previously with tert -butyl oxygen esters.

The thiol ester group 1 is the ester function of choice in condensation and acyl transfer reactions occurring in biochemical systems.' In contrast to (oxygen) esters or amides, relatively little is known about the electronic structure of this group. As a result of our interest in the chemistry and properties of thiol esters, we have undertaken a 13C NMR study of this class of compounds. A search of the literature has not produced any general NMR studies on the thiol ester 0

function.* We have thus obtained natural abundance 13C NMR spectra for some 30 different compounds. These results are discussed in connection with 13CNMR data available for other carbonyl derivatives.334 Substituent effects of the thiol ester group are analyzed and the effect of structure on thiol ester chemical shifts has been examined. Finally, we have focused attention on the relationship of these l 3 C NMR results to the chemistry and biological properties of the thiol ester function.

II

Experimental Section

1

Spectra. The I3C NMR spectra were obtained on ca. 20-25% (w/v) solutions in DCC13 or MezSO-d6 using a varian cFT-20 spectrometer

RCSR'

13C NMR Study of Thiol Esters

J.Org. Chem., Vol. 42, No. 12, 1977 2119

Table I. Carbon-13 Chemical Shifts for Thiol Ester Carbonyl Carbons0

0

It

RCSR' R' R

Methyl

Ethyl

Methyl

195.4

195.0 195.3 199.3

Ethyl

n-Propyl

Isopropyl

n-Butyl

195.0 195.7 199.2

194.9b 195.2

194.9 195.6

tertButyl

Isopropyl Cyclopropyl Phenyl

191.2

Benzyl

195.6 196.3 199.9 200.8 203.5 204.7 198.6

194.5 194.7

193.2 196.6

193.3

Phenyl 193.2 193.3

202.3

191.0 191.7

Chlorome thy1 1-Chloroethyl Dichloromethyl Acetyl

191.2 ( 18 9.2 )cpd

a 6, ppm from Me,, Si (internal standard) in Me, SO-d,. Numbers in italics refer to chemical shifts recorded in CDCl,. b The reported value' in dioxane is 1 9 4 . 1 ppm. CThe other carbonyl carbon in S-phenyl thiolpyruvate had a value of 193.2 ppm. d Registry no., 13884-99-6.

Table 11. Carbon-13 Alkyl Carbon Chemical Shifts for Thiol Ester Derivativesa

P f f

0

II

R

I

a'

p' Y'

6'

c-c-c-s-c-c-c-c

RC SR' Registry no.

0

R'

1534-08-3 625-60-5

Methyl Methyl

Methyl Ethyl

2307- 10-0

Methyl

Propyl

9 2 6 - 7 3-8

Methyl

Isopropyl

928-47-2

Methyl

Butyl

999-90-6

Methyl

tert-Butyl

32362-99-5

Methyl

Benzyl

934-87-2

Methyl

Phenyl

3232-39-1 2432-42-0 24 3 2-4 7-5 61540-13-4

Methyl Ethyl Ethyl Ethyl

Acetyl Ethyl Isopropyl tert-Butyl

29786-94-5

Isoprooyl

fer t-Bu tyl

61915-58-0 58058-56-3 1484-17-9 7269-35-4

Isopropyl Cyclopropyl Phenyl Phenyl

Benzyl tert-Butyl Ethyl Butyl

56377-45-8 56377-58-3 56377-47-0 61915-59-1

Chloromethyl Chloromethyl 1-Chloroethyl Dichloromethyl

tert-Butyl Benzyl tert-Butyl Butyl

P

CY

30.2 30.5b 30.5 30.5 30.6 30.6b 30.6 30.5 30.5 31.2 31.2 30.1 30.2 30.1 30.0 32.6 37.1 37.0 37.3 3 7.9 43.0b 43.4 42.4 22.4

48.1b 48.1 59.7 70.1

a' 11.3 23.1 23.6 30.5 23.0 34.4b 34.8 28.3C 28.9 47.4 4 7.8 32.7 33.4

9.5 9.4 9.2 9.6 19.2 19.4 19.1 9.6

21.5

22.7 34.1 47.1 47.5 46.9b 47.3 32.2 47.4 23.0 28.3 28.7 48.5b 33.0 48.8 (29.4)

0' 14.7 14.8 22.7 31.1 22.7 23.0 31.6 31.8 29.5 29.8

Y'

6'

13.1 13.3 21.6 22.1

13.5 13.6

21.7 22.1

13.5 13.6

21.3

13.4

14.8 22.8 29.4 29.9 29.6 29.9 29.6 14.7 31.4 31.8 29.3 29.5 (30.6)

a 6 p p m from Me,Si (internal standard) in Me,SO-d,. Chemical shifts recorded in italics refer t o values obtained in CDCl,. Assignments based in part o n off-resonance decoupled ' , C NMR spectra. CThe reported values' in dioxane are 30.1 ( a ) , 28.7 ( a ' ) ,32.1 ( p ' ) , 22.2 (y'),and 13.6 ( 6 ' ) .

b

with noise decoupling. The chemical shifts are referenced to internal Me&% The precision of the chemical shift data is at least h0.05 ppm (8K data points in the time domain for a 225 ppm spectral window). Materials. S -Phenyl thiolacetate, y-thiobutyrolactone, S-ethyl thiolpropionate, and diacetyl sulfide were obtained from Aldrich

Chemical Co. S-Methyl thiolacetate, S-isopropyl thiolacetate, Sisopropyl thiolpropionate, and S-ethyl thiolbenzoate were purchased from Wateree Chemical Co. S-Propyl thiolacetate and S-butyl thiolacetate were obtained from Columbia and S-ethyl thiolacetate and S-butyl thiolbenzoate were purchased from Pfaltz and Bauer. The purity of these commercial samples was checked by 'H NMR. They

2120

Hall and Wemple

J. Org. Chern., Vol. 42, No. 12, 1977

12I 2I 21

8 9 rl I

U *P r l l

"e:

I

f aJ

.-0 5

x X 0

.;; &

2 0

c

b

Y

.-09

2

were all used without further purification. S-tert-Butyl thiolacetate? S-tert -butyl S-benzyl thiolacetate? S-tert -butyl thioli~obutyrate,~ chlorothiolacetate,8 S-phenyl thiolpyruvate: S-tert-butyl cyclopropanecarbothioate,'O S-tert -butyl a-chlorothiolpropionate," Sbenzyl chlorothiolacetate," S-tert -butyl thiolpropionate,12 a methyl-y-thiobutyrolactone,lzS-tert-butyl P,B-diphenyl-P-hydroxythiolpropionate,'O S-benzyl b,b-diphenyl-P-hydroxythiolpropionate,1° S-phenyl (3,P-diphenyl-P-hydroxythiolpropionate,10 Stert-butyl a,a-dimethyl-P-phenyl-P-hydroxythiolpropionate,1o Sphenyl P-phenyl-P-hydroxythiolpropionate,l0and S-phenyl a-phenyl-~-hydroxythi~lpropionate~~ were prepared according to literature methods. S-Benzyl thiolisobutyrate was obtained from isobutyryl chloride, benzyl mercaptan, and pyridine according to the procedure described previously for the synthesis of S-tert -butyl bromothiolacetate." The S-benzyl thiolisobutyrate was isolated as a colorless oil following distillation under reduced pressure: bp 150-152 "C (15 m m ) ; . n z 3 ~ 1.5405; 'H NMR (CDC13,Me4Si) 6 1.18 (d, 6 H, J = 7 Hz), 2.74 (septet, 1 H, J = 7 Hz) 4.13 (s, 2 H), 7.33 (s, 5 H); IR (film) 1680 cm-l. Anal. Calcd for C11H140S:C, 68.00; H, 7.26; S, 16.50. Found: C, 68.31; H, 7.25; S, 16.27. S-Butyl dichlorothiolacetate was obtained in a similar way" from dichloroacetyl chloride, butyl mercaptan, and pyridine as a colorless oil: nZ8D1.4975; 'H NMR (CDC13, Me4Si) 6 0.75-1.15 (t, 3 H), 1.15-2.0 (m, 4 H), 3.03 (t, 2 H), 6.17 (s, 1H); IR (film) 1675 cm-'. Anal. Calcd for C6HloOSCIz: C, 35.83; H, 5.01; S, 15.94. Found: C, 35.66; H, 5.13; S, 16.12.

Results a n d Discussion The chemical shifts ( f i e , ppm from Me4Si) for a series of simple thiol esters are recorded in Tables 1-111. The signals corresponding to the carbonyl carbons as well as many of the carbons attached to halogen or oxygen heteroatoms are easily distinguished as a result of their low field chemical shifts. For other carbons, assignments were based on readily recognizable trends for a particular type of carbon observed within a series of similar structures. For example, eight different S-tert -butyl thiol esters were examined. In all of these spectra taken in M e 2 S o - d ~we observed a high intensity peak between 6 29.2 and 29.7 which was assigned to the methyl groups, and a low intensity peak between 6 46.6 and 48.5 which was assigned to the tertiary carbon of the tert-butyl group. Certain assignments were more difficult and are based solely on analogy with the corresponding carbon in the parent alkanes, alcohols, ethers, mercaptans, and/or oxygen esters reported previously.3g4 In many cases it was possible to confirm these assignments by single frequency off-resonance decoupling experiments. Where uncertainty still exists with respect to a given assignment, the number is indicated in brackets in Tables I-V. However, possible ambiguities in these assignments are not of major significance for the conclusions that we have drawn in our discussion. Of considerable interest is the result that the chemical shift of the carbonyl carbon in thiol esters generally occurs in the 193-203-ppm range for aliphatic thiol esters. This value is some 15-20 ppm further downfield than that found for the carbonyl carbon of all other carboxylic acid derivatives reported to date3,4 including the parent acids,I4J5 esters,14J6 amides,4J4J7 acid c h l ~ r i d e s , ~ aJ n~ hJ y~d r i d e ~ , 4 Jcarboxylate ~~ salts,15b and other derivative^.^ y-Thiobutyrolactone and a-methyl-y-thiobutyrolactone have carbonyl chemical shifts of 209.4 and 210.5 ppm, respectively, compared to a value of 178.0 ppm for y-butyrolactone.l4 This marked difference in chemical shift, characteristic of the thiol ester carbonyl carbon, may potentially be exploited in numerous ways including the use of the l3C NMR method in biochemical studies on coenzyme A thiol ester derivatives insofar as the thiol ester carbonyl carbon should appear further downfield than any other carbon in the complex coenzyme A structure. Our results would suggest a value of 193-195 ppm for the thiol ester carbonyl of acetyl COA.'~The 193-203-ppm range found for thiol esters closely approaches that reported for aldehydes and

13C NMR Study of Thiol Esters

J. Org. Chem., Vol. 42, No. 12, 1977 2121

Table IV. Carbon-13 Chemical Shifts for ,Simple Mercaptansa

Table V. Carbon-13 Chemical Shift Changes (A& ppm) Associated with Formation of Thiol Ester Derivatives of Mercaptans a

I'

Thiol ester Registry no. 107-03-9 75-33-2 109-79-5 75-66-1 100-53-8

Mercaptan Propyl mercaptanb Isopropyl mercaptan Butyl mercaptan C tert-Butyl mercaptan Benzyl mercaptan

p'

7'

(26.0)

(26.8)

12.8

S-tert -Butyl thiolacetate

29.9d

27.4d

23.6

35.6

21.0

13.4

S-Benzyl thiolacetate S-Isopropyl thiolpropionate S-tert-Butyl thiolpropionate

24.3 40.7 41.1 27.8

36.2 34.7 35.0

21.5

13.5

CY'

6'

S-Propyl thiolacetate S-Isopropyl thiolacetate S-Butyl thiolacetate

a 6 , .ppm from Me, Si (internal standard) in Me,SO-d,. Chemical shifts recorded in italics are values obtained in CDC1,. b The reported38values in CD, OD are 2 6 . 4 (CY'), 2 7 . 6 (p'), and 12.6 ('y'). CThe reported3*values in CD,OD are 24.6 (CY'),3 7 . 1 (by), 2 2 . 3 (y'), and 1 3 . 9 ( 6 ' ) . d h s i g n ments based in part on off-resonance decoupled ' C NMR spectra.

ketones.334 In this connection it is noteworthy that in many ways thiol esters resemble ketones or aldehydes in their chemical properties. For example, unlike oxygen esters, acids, or amides, thiol esters are rapidly reduced by sodium borohydride.20We have also recently found that the migratory aptitude of the thiol ester group is comparable to the ketone in the boron trifluoride induced rearrangement of a,P-epoxy carbonyl systems. Both groups migrate more readily than the oxygen ester.21 Many of the substituent effects observed earlier in 13C NMR studies of other carbonyl compounds were also found in our analysis of thiol esters. For example, as in the case of aldehydes, ketones, acids, esters, and amides,3p4 replacement of an a hydrogen with a methyl group causes a substantial downfield shift for the thiol ester carbonyl carbon (cf. Table I and the values for S. tert -butyl thiolacetate, thiolpropionate, and thiolisobutyrate) It is interesting that in the case of thiol esters the shift (-4 ppm) is somewhat greater than noted earlier (2-3 ppm) for aldehydes, ketones, acids, and oxygen esters.3 A relatively large increment (4.5 ppm) is found, however, in comparing acetamide with pr~pionamide.~b Substitution a t the a position with chlorine causes an upfield change (1-2 ppm) in the chemical shift of the carbonyl carbon in MezSO-d6 (cf. S-tert-butyl thiolacetate and chlorothiolacetate). A similar effect due to a-chlorine substitution has been observed for ketones, carboxylic acids, and acid chlorides.3~4Attachment of a phenyl group or a cyclopropane ring to the thiol ester carbonyl carbon also causes a substantial upfield shift (cf. Table I. Compare S-butyl thiolbenzoate with S-butyl thiolacetate and S-tert -butyl cyclopropanecarbothioate with S-tert-butyl thiolisobutyrate). In contrast to the marked changes in the carbonyl chemical shift caused by modification of the acyl portion of the thiol ester, relatively little change occurs when the hydrocarbon group attached to sulfur is modified. Thus, the carbonyl resonance for S-methyl, S-ethyl, S-propyl, S-isopropyl, S-butyl, and S-tert-butyl thiolacetates all come between 194.4 and 195.6 ppm in Me2sO-d~.An upfield shift of approximately 2 ppm occurs in S-phenyl thiolacetate (193.2 ppm in MezSO-d6) and the carbonyl carbon of diacetyl sulfide [(CH~CO)~S] comes at still higher field (191.8 ppm in CDCl3). A chemical shift of 194.5 ppm found for thioacetic acid has been ascribed to the thio-

S-tert -Butyl thiolisobutyrate S-Benzyl thiolisobutyrate S-tert -Butyl cyclopropanecarbothioate S-tert -Butyl chlorothiolacetate S-Benzyl chlorothiolacetate S-tert-Butyl achlorothiolpropionate S-Butyl thiolbenzoate

CY'

p'

y'

(+4.5) +4.4 +4.8 +4.6 +6.1 +6.7 +4.9 +4.2 +6.4 +6.5 +6.2 +6.3 +4.3 4-6.7

(-4.1) -4.7 -4.1 -4.4 -5.2 -5.2

+0.3

+7.8

-5.3

+5.1 +7.5

-5.2

+4.7 +4.4

-4.2 -4.5

6'

+0.6 +0.1 +0.5 +0.1

-4.6 -5.3 -5.1 -5.1 -5.1 -5.1

+0.6 +0.6

f0.1 +0.1

a The data represent chemical shifts of the S-alkyl carbons of the indicated thiol ester relative to the analogous carbon in the corresponding mercaptans. Comparisonswere made in MezSO-dG or CDC13 (italics).The values were calculated from data in Tables I1 and IV.

carbonyl group in tautomer Z4 However, the weight of experimental evidence argues against a thiocarbonyl group in thioacetic acid favoring instead a carbonyl group as in tautomer 3,1c922The 194.5-ppm value obtained for thioacetic acid is in good agreement with structure 3 in view of the result that S

II CH,COH 2

0

1I

CHSCSH 3

S-alkyl and 5'-aryl thiolacetates all fall in the 193-196-ppm range. A semiempirical approach has been developed to relate carbonyl chemical shifts to the A bond polarity of the carbonyl function.23 This A bond polarity depends to a considerable extent on the relative electron-withdrawing ability of attached groups as gauged by Taft's inductive parameter ~ 7 1 . 2Based ~ on this analysis the lower electron-attracting ability of sulfur compared to oxygen would result in greater shielding of the carbonyl carbon in oxygen esters relative to the corresponding thiol esters.14J6 The effect of p-p ir bonding in thiol esters on the chemical shift of the carbonyl carbon is less certain. Indeed the relative degree of resonance in thiol esters has recently been questioned by NO^^^ in a DNMR study of rotational barriers in thioacetic acid. This work suggests that there is considerably more resonance in thio acids than in the corresponding oxygen acids or esters in contradiction to conclusions reached in earlier s t ~ d i e s . ~ c , 2 6 ~ ~ ~ The deshielding effect of the thiol ester function a t the (Y carbon is very similar to that caused by a ketone or aldehyde group. Thus, the methyl carbon in S-alkyl and S-aryl thiolacetates occurs in the 30-31-ppm range (Table 11) while the methyl carbon in acetaldehyde comes at 31 ppm and in aliphatic methyl ketones between 28 and 30 ~ p mIn. contrast, ~ the methyl carbon in acetate e ~ t e r s or ~ acetamide~l7~ J~~ occurs a t about 21 ppm.3J6fIn this connection it is interesting to note

2122 J . Org. Chem., Vol. 42, No. 12, 1977 that the acidity of the ai protons in thiol esters is comparable to that in ketones while both ketones and thiol esters are substantially more acidic than the corresponding oxygen esFurther, we have found recently that the nucleophilicity of thio: ester lithium enolates in substitution reactions with alkyl halides is considerably less than that of the corresponding oxygen ester lithium enolates.12 Thus again in this instance we see a useful correlation of the 13CNMR data with the chemical properties of thiol esters as compared to data available for ketones, esters, and amides. For simple thiol esters such as S-ethyl, S-phenyl, or S-butyl thiolacetates there is relatively little effect on chemical shift values as a result of a change in solvent from the very polar Me2SO to the less polar chloroform. Generally we observed a slight downfield shift of between 0.1 and 0.6 ppm on going from MezSO-d6 to CDC13 for the carbonyl carbon as well as most of the other carbons. This is consistent with earlier observations on the small effect of a change in solvent on the carbonyl chemical shift of acetone.29In contrast in a study of several @-hydroxythiol inters we observed a marked downfield shift of the carbonyl carbon of between 1.8 and 5.0 ppm on going from M e 2 S o - d ~to CDC13 (Table 111).It is likely that intramolecular hydrogen bonding is responsible for this solvent effect. Intramolecular hydrogen bonding would be relatively important in chloroform but less so in Me2S0, where intermolecular hydrogen bonding with solvent molecules is expected to be more significant. Me2SO has recently been used to rupture intramolecular hydrogen bonds in NMR studies of @-hydroxy (oxygen) esters and amides.30~31Hydrogen bonding interactions with ketone and ester carbonyl functions is reported to result in a downfield shift of the carbonyl carbon atom.14a,29,32Additional support for strong intramolecular hydrogen bonding in @-hydroxythiol esters in CDC13 is seen in the relatively large chemical shift difference of the two amethyl carbons in S-tert -butyl a,a-dimethyl-@-phenyl-@hydroxythiolpropionate (4). The methyls occur at 6 20.0 and 21.5 in Me2sO-d~.One is shifted upfield (6 19.5) and the other downfield (6 23.5) in CDC13. Intramolecular hydrogen bonding in CDC13 would result in restricted rotation about the C,-Cp and Cc=o-C, bonds in 4 such that one methyl would be placed in close proximity to the phenyl group.

Hall and Wemple ters are in the same direction although somewhat larger than those found for oxygen esters.16fshThe fact that relatively little variation is seen in the PAS values (5.2 f 0.2 in MezSO-d6) for a large number of S-tert-butyl esters would indicate that the C-2 esterification effect results primarily from interaction of the S-tert-butyl group with the carbonyl function rather than the R substituent in the acyl group. A similar pattern is seen for other types of S-alkyl thiol esters studied (Table V). This would support the conclusion that 2 conformation 5 is the major conformation present in these thiol esters.

Z 5

E 6

The origin of nonequivalence in the chemical shifts of syn and anti methyl groups in N,N-dimethylformamide has been attributed to electric field effects33 as well as steric perturb a t i o n ~ . l ~An a *electric ~~ field argument has been suggested to account for the large a‘A6 value (C-1 esterification effect) found for tert-butyl formate compared to a smaller value obtained for other formate esters.16fMore recently chemical shifts have been evaluated for a large number of methyl, ethyl, isopropyl, and tert-butyl oxygen esters.lGhA linear relationship was shown to exist between the 13Cchemical shifts of the a’ carbon and the pK, values of acids from which the esters were derived. This was explained as a consequence of the polar character of the -C,,6+-026-CR bond. With respect to variation of the 0-alkyl group from primary to secondary and tertiary, the C-1 esterification effect (a’A6 values) can be correlated with increasing stability of the partial positive charge a t the a’ carbon.16h These results16h suggest that the electric field argument is not a major factor in determining this C-1 esterification effect. The results obtained with oxygen esters16f,hdo not appear to be inconsistent with steric perturbation providing some contribution to the large a’A6 values found for tert -butyl oxygen esters. Steric perturbation resulting in greater deviation from ~ o p l a n a r i t y 3 ~ for3 ~the ~ ester function has recently been proposed to account for the large bathochromic shift in the ultraviolet found for tert -butyl 0/H*..() acetate compared to other alkyl acetates.37 The a’A6 values for thiol esters (Table V) increase with increasing acidity of the acid from which the ester is derived in agreement with the conclusions of Pelletier.lGh The C-1 esterification effect found for propyl, butyl, or isopropyl thiol “CH3 esters is in the same direction and generally larger than that CH3 found for propyl, butyl, and isopropyl oxygen esters; however, 4 the a’A6 value found for S-tert-butyl thiol esters is considerably smaller than that found for tert-butyl oxygen esters. Also of interest is the result that the a carbon in the pThe relatively small difference (-2 ppm) in a‘A6 values behydroxy thiol ester system is shielded (0.1-0.7 ppm) while the tween S-tert-butyl thiol esters and other types of S-alkyl thiol p carbon is deshielded (1.1-2.2 ppm) as a result of intramoesters is of particular interest and should be contrasted with lecular hydrogen bonding in CDC13. This phenomenon was also seen in our analysis of 4-hydroxy-4-methylpentan-2-one the very large difference (-10 ppm) found earlier for oxygen esters.16zh In the thiol ester system we may expect some in[CH&OCH&(OH)(CH3)2], which gave chemical shifts of crease in the degree of polarization of the S-C,’ bond when 208.3 and 210.5 ppm for carbon 2,55.8 and 54.2 ppm for carcomparing S-tert-butyl esters with other alkyl thiol esters, bon 3, and 68.6 and 69.6 ppm for carbon 4 in M e ~ S 0 - dand ~ although this effect is expected to be smaller than in oxygen CDC13, respectively. Shielding of the a position is then due esters1Gh owing to the lower electronegativity of sulfur comto increased importance of a fixed conformational state that pared to oxygen. It would seem that this X-C,’ polarization is characteristic of the intramolecularly hydrogen bonded exp1anationlGhis sufficient to account for the variation in a’6 structure in CDC13 as opposed to the greater variety of convalues found for thiol esters without invoking steric perturformations available tlo the p-hydroxy thiol ester in Me2SObation in the S-tert-butyl thiol ester system. The larger size d6. of sulfur compared to oxygen would leave the tert-butyl furWe have obtained 13C NMR spectra of several simple ther removed from the acyl group in S-tert -butyl thiol esters mercaptans (Table IV) and with this information calculated relative to tert-butyl oxygen esters. The likelihood of steric a’ and FA6 values for certain thiol ester derivatives (Table V). perturbation is greater in the tert-butyl oxygen ester system. The FA6 values (C-2 esterification effect) found for thiol es-

A

J . Org. Chem., Vol. 42, No. 12, 1977 2123

5-Thiacyclohexenecarboxaldehydes

To what extent this influences the a'A6 values in tert-butyl oxygen esters is not clear based on available information.

Acknowledgment.This research was sL2ported in part by a grant from the U.S. Public Health Service (Research Grant R01-CA17719-02). R e g i s t r y No.-Isobutyryl

chloride, 79-30- 1; dichloroacetyl chlo-

ride, 79-36-7.

References and Notes (1) For reviews of thiol ester chemistry see (a) T. C. Bruice, "Organic Sulfur Compounds", N. Kharasch. Ed., Pergamon Press, Oxford, 1961, p 421; (b) T. C. Bruice and S.J. Benkovic, "Bioorganic Mechanisms", Vol. 1, W. A. Benjamin, New York, N.Y., 1966, p 259; (c) M. 3. Janssen, "The Chemistry of Carboxylic Acids and Esters", S. Patai, Ed., Wiley-lnterscience, New York, N.Y., 1969, p 705. (2) The 13CNMR spectra of several diethyl thlolmalonates have recently been reported: R. Radeglia and S.Scheithauer, 2.Chem., 14, 20 (1974). The spectrum of S-butyl thlolacetate is recorded in the collection of spectra compiled by L. F. Johnson and W. C. Jankowski. "Carbon-13 NMR Spectra", Wiley-lnsterscience, New York, N.Y., 1972. (3) J. B. Stothers, "Carbon-13 NMR Spectroscopy", Academic Press, New York, N.Y., 1972. (4) G. C. Levy and G. L. Nelson, "Carbon-13 Nuclear Magnetic Resonancefor Organic Chemists", Wiley-Interscience, New York, N.Y., 1972. (5) A. A. Schleppnik and F. B. Zienty, J. Org. Chem., 29, 1910 (1964). (6) B. K. Morse and D. S.Tarbell, J. Am. Chem. SOC., 74, 416 (1952). (7) P. J. Lillford and D.P. N. Satchell, J. Chem. Soc. 6, 1303 (1970). (8) T. P. Dawson, J. Am. Chem. Soc., 69, 1211 (1947). (9) R. A. Gorski, D.J. Dagli, V. A. Patronlk. and J. Wemple, Synthesis, 611 (1974). (10) J. Wemple. Tetrahedron Lett., 3255 (1975). (11) D.J. Dagli, P. S. Yu, and J. Wemple, J. Org. Chem., 40, 3173(1975), (12) R . A. Groski, G. Wlolber, and J. Wemple, Tetrahedron Lett., 2577 (1976). (13) J. Domagala and J. Wemple, Tetrahedron Lett., 1179 (1973). (14) (a) J. B. Stothers and P. C. Lauterbur, Can. J. Chem., 42, 1563 (1964); (b) G. A. Gray, P. D. Ellis, D. D. Traficante, and G. E. Maciel, J. Mgn. Reson., 1, 41 (1969). (15) (a)E.LippmaaandT. F'ehk, Kem. Teollisuus, 24, 1001 (1967);(b)R.&gen and J. D. Roberts, J. Am. Chem. Soc.,91,4504 (1969); (c) H. Brouwer and J. B. Stothers, Can. J. Chem., 50, 601 (1972). (16) (a) K. S.Dhami and J B. Stothers, Can. J. Chem., 45, 233 (1967); (b) D.

H. Man and J. B. Stothers, ibid, 45,225 (1967); (c) E. Lippmae and T. Pehk, EestiNSV Tead. Akad. Toim., Keem., Gsol., 17,210 (1968); (d)M. Christl, H. J. Reich, and J. D.Roberts, J. Am. Chem. Soc., 93,3463 (1971); (e) M. Gordon, S. H. Grover, and J. B. Stothers, Can. J. Chem., 51,2092 (1973); (f) D. E. Dorrnan, D. Bauer, and J. D. Roberts, J. Org. Chem., 40, 3729 (1975); (g) A. C. Rojas and J. K. Crandall, ibid., 40,2225 (1975); (h) S. W. Pelletier, F. Djarmati, and C. Pape, Tefrahedmn, 32, 995 (1976). (a) G. C. Levy and G. L. Nelson, J. Am. Chem. Soc.,94, 4897 (1972); (b) D.E. Dorman and F. A. Bovey. J. Org. Chem., 38, 1719 (1973). (a) G. Mlyazima, Y. Utsumi, and K. Takahashi. J. Phys. Chem., 73, 1370 (1969); (b) K. Frei and H. J. Bernstein, J. Chem. Phys., 38, 1216 (1963). NMR studies on coenzyme A and derivatives have been reported. See C.4. Lee and R. H. Sarma, J. Am. Chem. Soc.,97, 1225 (1975). E. J. Barron and L. A. Mooney, Anal. Chem., 40, 1742 (1968). D.J. Dagli, R. A. Gorski, and J. Wemple, J. Org. Chem., 40, 1741 (1975), and references cited therein. (22) R. B. De Alencastro and C. Sandorfy, Can. J. Chem., 51, 1443 (1973);W. H. Hocking and G. Winnewisser. J. Chem. Soc., Chem. Commun., 63 (1975); Z.Naturforsch. A, 31, 422, 438 (1976). (23) G. E. Maciel, J. Chem. Phys., 42, 2746 (1965). (24) R. W. Taft, E. Price, I. R. Fox, I.C. Lewis, K. K. Anderson, and G. T. Davis, J. Am. Chem. Soc., 85, 709 (1963), and references cited therein. (25) E. A. Noe, J. Am. Chem. Soc., 99,2803 (1977). We are grateful to Dr. Noe for supplying us with a reprint of this work prior to publication. (26) For a review see C. C. Price and S.Oae, "Sulfur Bonding", Ronald Press, New York, N.Y., 1962. (27) I. Wadso, Act8 Chem. SCend., 18, 487 (1962). (28) L. Wessely and F. Lynen, fed. hoc., Fed. Am. SOC.Exp. Biol., 12, 658 (1953); C. Schwarzenbach and E. Felder, Helv. Chim. Acta, 27, 1701 (1944). (29) G. E. Maciel and G. C. Ruben, J. Am. Chem. SOC., 85,3903 (1963). The carbonyl chemical shifts of esters are somewhat less sensitive to changes in solvent: G. E. Maciel and J. J. Natterstad, J. Chem. Phys., 42, 2752 (1965). (30) H. 0. Kalinowski, B. Renger, and D. Seebach,Angew. Chem., ht. Ed. Engl., 15, 234 (1976). (31) M. Branik and H. Kessler, Tetrahedron, 30, 781 (1974). (32) P. C. Lauterbur, Ann. N. Y. Acad. Sci., 70,841 (1958); G. E. Maciel and G. 8. Savitsky, J. Phys. Chem., 88, 437 (1964). (33) W. McFarlane, J. Chem. SOC., Chem. Commun., 418 (1970). (34) N. K. Wilson and J. B. Stothers. Top. Stereochem., 8, 1 (1974). (35) A. G. Pinkus and E. Y. Lin, J. Mol. Struct., 24, 9 (1975). (36) G. I. L. Jones and N. L. Owen, J. Mol. Struct., 18, 1 (1973), and references cited therein. (37) L. R . Caswell, M. F. Howard, and T. M. Onisto, J. Org. Chem., 41, 3312 (1976). (38) G.Barbarella, P. Dembech. A. Garbesi, and A. Fava, Org. Magn. Reson., 8, 108 (1976).

Phase-Transfer Catalyzed Syntheses. 5-Thiacyclohexenecarboxaldehydes and 3,4-Epoxy-2,5-dihydrothiophenes' John M. McIntosh* and Hamdy Khali12 Department of Chemistry, University of Windsor, Windsor, Ontario, Canada N9B 3P4 Received October 13,1976 The phase-transfer catalyzed condensation of 3-thioacetoxyaldehydes w i t h acrolein and crotonaldehyde leads t o cyclized products 5-9. Product distributions indicate that no equilibration of intermediates occurs as has been previously noted in pyridine solution. Condensations of a-mercaptocarbonyl compounds w i t h 2-chloroacrylonitrile led t o epoxides 11-14 in excellent yields. This reaction may be of some importance in biotin synthesis.

Recently, we have :shownthat the reaction of 3-mercaptoaldehydes (1) with conjugated carbonyl compounds affords, by a conjugate addition-aldol condensation sequence, an excellent route to substituted 5-thiacyclohexene-1-carboxaldehydes Compounds related to 2 have previously been shown4 to be excellent synthons for stereospecific alkene synthesis. However, )two drawbacks to the synthesis, as reported,3 are evident. The first is the difficulty encountered in the purification of mercaptans 1. Whereas the isomeric 2-mercaptoaldehydes exist largely as dimeric 2,5-dihydroxy- 1,4-dithianes5which can be purified relatively easily, 1 are polymeric hemithioacetals which are uniformly evilsmelling, viscous oils that decompose (presumbly by dehydration) when distillation is attempted. This leads to com-

1

2

plicated mixtures when the preparation of 2 is attempted. Although many examples of 1 are obtained pure enough for direct use in the cyclization, others are not and it was felt that an alternate preparation which avoided this difficulty would be desirable. Replacement of the thiol proton with a suitable protecting group which could be converted into the anion of 1 in situ would achieve this end. Furthermore, as the malodorous properties of most thiols are associated with the SH